Вы находитесь на странице: 1из 14

Heat release model for the combustion of diesel oil

emulsions in DI diesel engines


Jamil Ghojel
a,
*
, Damon Honnery
b,1
a
Department of Mechanical Engineering, Monash University, Cauleld Campus,
P.O. Box 197, Cauleld East 3145, Australia
b
Department of Mechanical Engineering, Monash University, P.O. Box 31, Clayton Campus, Vic. 3800, Australia
Received 3 November 2004; accepted 30 January 2005
Abstract
Diagnosing the combustion process in internal combustion engines using cylinder pressure, fuel delivery
pressure and injector needle lift data is a well-established and widely used procedure in engine and fuel per-
formance tests. Of the models developed, the rst law single zone is the simplest and easiest tool to use for
quick preliminary analysis of engine performance. It can yield valuable information about the eect of
engine design changes, fuel injection system, fuel type and engine operating conditions on the combustion
process and engine performance. The model described in this paper is based on a model designed for both
SI and CI engines burning liquid and gaseous hydrocarbon fuels. In its current form, the model can be used
for diesel engines burning water/diesel emulsions and allows for the eect of the presence of inert water in
the combustion products on the ratio of specic heats and heating value of the fuel. The output from the
model includes the apparent heat release, fraction of fuel burned, fuel burning rate, heat losses, indicated
parameters and average gas temperature. Some results of analysis of the performance of a DI diesel engine
burning a standard diesel fuel (referred to in the texts as certied diesel fuel CDF, or simply diesel fuel) and
diesel oil emulsion (DOE) are presented.
2005 Elsevier Ltd. All rights reserved.
Keywords: Heat release; Emulsion; Fuel; Diesel; Engine
1359-4311/$ - see front matter 2005 Elsevier Ltd. All rights reserved.
doi:10.1016/j.applthermaleng.2005.01.016
*
Corresponding author. Tel.: +61 3 9903 2490; fax: +61 3 9903 2766.
E-mail addresses: jamil.ghojel@eng.monash.edu.au (J. Ghojel), damon.honnery@eng.monash.edu.au (D. Honnery).
1
Tel.: +61 3 9905 1988.
www.elsevier.com/locate/apthermeng
Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
Nomenclature
c
p
constant pressure specic heat (kJ/kg K)
c
v
constant volume specic heat (kJ/kg K)
h
c
convective heat transfer coecient (W/m
2
K)
h
r
radiative heat transfer coecient (W/m
2
K)
LHV lower heating value of the fuel (MJ/kg)
M molar composition per mass of fuel (kmol/kg fuel)
M
a
airfuel charge (kmol/kg fuel)
Nu Nusselt number
n mean polytropic index
n
0
intermediate polytropic index
N
1
increments from close of the inlet valve to start of fuel injection
N
2
increments from the start of combustion to the exhaust valve opening
p pressure (bar)
p
r
reference pressure (bar)
q heat release (kJ/kg mixture)
R ratio of crank radius to connecting rod length
R universal gas constant (kJ/kmol K)
Re Reynolds number
r compression ratio
T temperature (K)
T
g
cylinder gas temperature (K)
T
w
cylinder wall temperature (K)
T
0
ambient temperature (K)
T
r
residual exhaust gas temperature (K)
V volume (m
3
)
V
cl
cylinder clearance volume (m
3
)
v specic volume (m
3
/kg)
x mass fraction
Subscripts
a at the start of compression
c at the start of combustion, carbon
h hydrogen
i position or event
inj at the start of fuel injection
m measured
w water
Greek symbols
a residual gas coecient
b corrected measured pressure ratio
2 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
1. Introduction
The direct injection (DI) diesel engine, both naturally aspirated and turbocharged is a highly
ecient heat engine suitable for use in road, rail and sea transport applications and also for
numerous stationary applications. Signicant progress has been made in the last decade in further
improving the thermal eciency of the diesel engine and reducing soot and noise levels thanks
largely to the introduction of common rail fuel injection system and advanced electronic injection
controls. However, the dual problem of NO
x
and particulate emissions has persisted due to the
contradictory requirements their for suppression. The common rail fuel injection system enhances
the mixing of fuel and air and increases the combustion temperature at the same time. The former
reduces particulates emissions and the latter causes the emissions of NO
x
to increasehence the
need for particulateNO
x
tradeo strategies in the engine control system. Research to date [111]
has shown that using diesel/water emulsion as an alternative fuel to standard diesel fuel can sig-
nicantly reduce the emission levels of both pollutants if the penalty in reduced power can be tol-
erated. This can be achieved without any need to modify or retrot the engine.
With the increased interest in the diesel oil emulsion (DOE) as an alternative fuel for diesel en-
gines, there is a need for the development of relatively simple and fast engine combustion diagnos-
tic tool. Many models have been developed to determine the burning rate of the fuel in internal
combustion engines (ICE). The single-zone model, as compared with multi-zone models [12,13], is
much simpler, quicker and easier to run. The accuracy of the heat release model is largely depen-
dent on the accuracy of the measured pressure data, accuracy of locating the top dead centre and
v apparent heat release fraction
DH
0
lower heating value of the fuel per unit mass of the mixture (kJ/kg mixture)
/ equivalence ratio (stoichiometric airfuel ratio/actual airfuel ratio)
c
p
ratio of specic heats c
p
/c
v
of combustion products
k relative airfuel ratio (1//)
l
f
fuel molecular weight (kg/kmol)
h crank angle (degrees)
n coecient of combustion eectiveness
w cylinder volume to cylinder clearance volume ratio
Abbreviations
BDC bottom dead centre
CDF certied diesel fuel
CI compression ignition
DI direct injection
DOE diesel oil emulsion
ICE internal combustion engine
TDC top dead centre
NO
x
oxides of nitrogen
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 3
ARTICLE IN PRESS
the computed values of the ratio of specic heats of the products c
p
[14]. The accuracy of the latter
can be improved by assuming complex multi-species equilibrium composition of the products of
combustion; however, it is well known that the solution of the many thousands of complex equi-
librium composition reactions, typically required in the analysis of ICE combustion, is both
lengthy and costly in terms of computer time [15]. A previous study [16] demonstrated that the
much simpler xed (frozen) composition analysis based on the assumption of complete combus-
tion yields reasonably accurate results of c
p
within the ranges of temperature T = 8002500 K,
pressure p = 1100 bar and relative airfuel ratio k = 0.73.35 (equivalence ratio / = 0.31.43).
Heat released in the cylinder of an internal combustion engine as a result of fuel combustion is
usedto heat the working uid(increase its internal energy) anddo external work. Part of the released
heat is also lost to the cooling mediumthrough the walls of the cylinder and cylinder head during the
combustion and expansion processes. Small fraction of the heat energy of the fuel is also lost as a
result of dissociation of some combustion products and incomplete chemical reactions.
2. Heat release model
The main input to the model is the measured trace of the in-cylinder pressure versus crank angle
from the test engine. The test engine used for these measurements is a fully instrumented 4-cylin-
der, 4-stroke, water-cooled, direct injection industrial Hino diesel engine, Fig. 1. One cylinder is
monitored for gas pressure, injector needle lift and fuel delivery pressure. The crank angle incre-
ment at which pressure is measured in the current investigation is 0.2. The true TDC, reference
pressure for the measured data and polytropic compression index are calculated thermodynami-
cally. The model is a single-zone thermodynamic model, which assumes ideal gas behaviour for
the airfuel mixture. At any position of the crank angle the control volume formed by the cylinder







70 litre fuel tank
Exhaust
Inlet charge air
In-cylinder conditions:
Pressure, needle lift and
fuel injection pressure
HINO ENGINE
Test cell
ambient conditions
Measurement
fuel tank (12 litre)
Eddy current
dynamometer
Torque load
cell
Engine monitoring PC

Combustion monitoring PC Exhaust monitoring PC
Exhaust gas emission
measurement
Temperature loom:
Engine oil, fuel, cooling water,
inlet air and exhaust.
RPM and crank
angle sensor
Fuel return
Fuel mass flow
load cell
Inlet air flow rate
sensor
Fig. 1. Experimental setup of the test engine. The engine is a 4-cylinder, 4 stroke, direct injection diesel with bore
104 mm, stroke 118 mm, displacement 4.009 l and compression ratio 17.9.
4 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
and cylinder head walls contains homogeneous air and products of combustion depending on the
airfuel ratio. The presence of residual gases is accounted for by the coecient of residual gases.
The change in volume as a result of combustion is accounted for by the coecient of molecular
change. The fuel delivery and burning rates are assumed equal and the model uses the estimated
trapped mass of airfuel mixture in the combustion chamber to calculate the heat release rate of
the fuel. The thermodynamic property c
p
is calculated as a function of the temperature and pres-
sure of the trapped combustion products, overall relative airfuel ratio with dissociation ignored.
The temperature computed is the average temperature for the trapped mass and is a function of
crank angle only. The following sections summarise the dierent components of the model.
2.1. The compression process
Fig. 2 shows the experimental pressure trace during compression and expansion in the moni-
tored cylinders at 200 N m and 2200 rpm with the computed motoring pressure. The compression
process in real engines is polytropic starting from the moment the inlet valve closes and ending
when the injection process starts. The mean polytropic index n of the compression process is cal-
culated for this range using an equation based on the least square method for N
1
pressure values
between the above mentioned range and selected at 0.2 crank angle increments [17],
n

N
1
i1
ln
p
inj
p
i
ln
W h
i

W h
inj
_ _

N
1
i1
ln
W h
i

W h
inj
_ _
2
: 1
The function W(h
i
) is the ratio of the cylinder volume at any point to the clearance volume V
cl
,
Fig. 2. Firing and motoring pressure proles at a load of 200 N m and speed 2200 rpm.
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 5
ARTICLE IN PRESS
W h
i

V
i
V
cl
1
r 1
2
1
1
R
_ _
cos h
i

1
R

1 R
2
sin
2
h
i
_ _ _ _ _
: 2
The pressure is the absolute pressure and is taken as the sum of the measured pressure p
m
and a
reference pressure p
r
. The latter is calculated using a technique based on the method presented in
Ref. [18]. The equation of the polytropic process pv
n
0
= constant can be written for two values of
the crank angle position (h
1
and h
2
) within a narrow range during the compression stroke as
follows:
p
m
1
p
r
p
m
2
p
r

W h
2

W h
1

_ _
n
0
b; 3
from which,
p
r

bp
m
2
p
m
1
1 b
: 4
A value of n
0
in Eq. (3) is assumed and p
r
is calculated from Eq. (4). This value of p
r
is added to
the measured pressures p
inj
and p
i
and n is calculated from Eq. (1). The hot motoring pressure is
now calculated for the crank angle range h = 180540 (compressionexpansion) from Eq. (5)
below,
p
i

p
inj
W h
i

W h
inj
_ _
n
: 5
The value of n
0
is adjusted until n
0
% n and the area between the proles of the corrected mea-
sured pressure (p
m
at location the crank angle position plus p
r
) and the calculated pressure (Eq.
(5)) is minimized. The TDC is then adjusted by shifting the entire corrected pressure trace to the
right or to the left until the measured and calculated values are further reconciled.
If we assume that the compression process starts at bottom dead centre (BDC), the pressure at
that point (p
a
) can be determined by putting h
i
= 180 in Eq. (3), which yields
p
a

p
inj
r
W h
inj
_ _
n
: 6
The temperature at the start of the compression process is given by
T
a

T
0
DT aT
r
1 a
; 7
where the coecient of residual gases a % 0.04 for diesel engines and DT is the temperature rise of
the fresh charge from contact with the walls of the inlet manifold. The temperature at the start of
combustion is given by
T
c
T
a
r
W h
c

_ _
n1
: 8
6 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
2.2. Ratio of specic heats
The ratio of specic heats for the combustion products c
p
of any fuel can be calculated if the
composition and average temperature of the products are known. The volumetric composition
of the DOE fuel used in this investigation is 13% H
2
O, 85% diesel fuel and 2% additives (surfac-
tant and cetane improver). This gives a molecular mass l
f
= 143.5 if the diesel fuel is assumed to
be represented by the hydrocarbon C
12
H
21.5
. The composition by mass is then 15.06%, 82.72%
and 2.22% for water, diesel fuel and additives, respectively. The stoichiometric airfuel ratio,
quantity of airfuel charge and the fractions of the products of complete combustion of a lean
mixture based on this composition can be written as follows:
A
F
_ _
stoich

1
0:21
x
c
12

x
h
4
_ _
; 9
M
a
k
A
F
_ _
stoich

1
l
f
; 10
M
CO
2

x
c
12
; M
H
2
O

x
h
2
x
w
; M
O
2
0:21k 1
A
F
_ _
stoich
;
M
N
2
0:79k
A
F
_ _
stoich
: 11
In these equations, x
c
, x
h
and x
w
are the mass fractions of carbon, hydrogen and water in one
kilogram of fuel (0.7197, 0.1075 and 0.1506, respectively). The lower heating value of the emulsi-
ed fuel can be estimated using the composition of the combustion products given by Eq. (11)
when k = 1 (stoichiometric mixture) if the enthalpy of formation of the fuel is known. Alterna-
tively, it can be estimated as follows [19]:
LHV 34:013x
c
125:6x
h
2:512 9x
h
x
w
: 12
Eq. (12) yields a value of 35.2 MJ/kg, which is about 1.3% lower than the value determined
experimentally for this fuel (35.7 MJ/kg).
The ratio of specic heats is calculated for the reference diesel fuel and diesel oil emulsion from
the constant-pressure specic heats of the products of complete combustion CO
2
, H
2
O, O
2
and N
2
[20] and their mole fractions in the products. The computed values of c
p
are shown in Fig. 3 as a
function of the relative airfuel ratio k at three average gas temperatures: 1000, 2000 and 3000 K.
The diesel fuel exhibits consistently higher c
p
than the diesel oil emulsion with the dierence
increasing directly with the temperature and inversely with the relative airfuel ratio k. The values
of c
p
for the diesel fuel are higher than for the DOE over the shown range of k by 1.52%, 3.3
6.3% and 4.28.4% at 1000, 2000 and 3000 K, respectively. These seemingly small variations in c
p
can have appreciable eect on the calculated heat release characteristics.
2.3. Apparent heat release
The combustion process is assumed to start from the moment the gas pressure separates from
the polytropic compression process. If we divide the process into small equal increments (0.1,
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 7
ARTICLE IN PRESS
0.2, 0.5 or 1.0 crank angle), the specic heat released over the increment (i + 1) i in units of J/
(kg mixture) will be equal to the energy required to raise the internal energy of the gas plus the
useful work done over the same increment,
q
i1i
c
vi1i
T
i1
T
i

_
v
i1
v
i
pdv; 13
where c
v(i + 1) i
is the mean specic heat at constant volume over increment from i to (i + 1), and
T
i
and T
i + 1
are the absolute temperatures of the gas at the start and end of increment. This equa-
tion can be written in terms of pressure, volume and ratio of specic heats as
q
i1i

v
a
r
p
i1
p
i
2
W h
i1
W h
i

p
i1
W h
i1
p
i
W h
i

c
i1i
1
_ _ _ _
: 14
The specic cumulative heat release can now be written as
q
i

N
2
i1
q
i1i
; 15
where N
2
in Eq. (15) is the number of increments of the combustion process from the start of the
process to the moment the exhaust valve opens. The specic volume of the products can be deter-
mined from the equation of state,
v
a

RT
a
M
a
p
a
: 16
The fraction cumulative apparent heat release is given by
v
i

q
i
DH
0
17
Fig. 3. Ratio of specic heats for the diesel and emulsied fuels as a function of the relative airfuel ratio and average
mixture temperature.
8 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
and the rate of heat release by
dv
i
dh

v
i1
v
i
Dh
; 18
where Dh = 0.2 crank angle in the current study. DH
0
is the lower heating value per unit mass of
mixture (kJ/kg mixture). At the end of the heat release process,
v
i max

q
i max
DH
0
n:
The parameter n is called the coecient of combustion eectiveness and denotes the maximum
value of the apparent heat release. The temperature T
i
of the products of combustion is computed
using the equation of state from the start of combustion till the opening of the exhaust valve:
T
i

T
c
p
c
W h
i

W h
c

p
i
: 19
2.4. Heat release characteristics
Fig. 4 shows typical heat release characteristics of the test engine operating on the standard die-
sel fuel. The apparent cumulative heat release (fraction) and apparent rate of heat release (1/s) are
calculated from the measured pressure data according to the method described above. These char-
acteristics are sucient to assess the performance of an engine, eect of operating conditions on
engine performance or compare the performance of dierent engines under the same operating
conditions. In the current study, the eect of changing the fuel from CDF to DOE is given
as an example of the models capability. The maximum value of the cumulative heat release
is reached at crank angle 460 for this particular test condition. The two main phases of the
Fig. 4. Typical heat release characteristics of the test engine operating on the standard diesel fuel.
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 9
ARTICLE IN PRESS
combustion process, premixed and diusion, are clearly seen in the rate of heat release curve. If all
heat losses (due to heat transfer from the gases to the cylinder walls, dissociation, incomplete
combustion, gas leakage) are added to the apparent heat release characteristics, the fuel burn
characteristics are obtained. The latter are shown in Fig. 4 in bold lines and the cumulative curve
approaches unity at the end of the combustion process. The method used to estimate heat losses is
explained in the following section.
2.5. Heat losses
The apparent heat release characteristics (cumulative heat release and rate) calculated using
Eqs. (17) and (18) relate to the heat released to raise the internal energy of the gases and produce
mechanical work. The dierence between the heat energy available in the fuel and the apparent
heat release represents all losses including heat transfer from the gases to the engine cylinder walls,
incomplete combustion, gas leakage and dissociation. The heat losses due to incomplete combus-
tion, gas leakage and dissociation are usually ignored and heat transfer losses by convection and
radiation are estimated using empirical correlations. In most cases the radiation component is ig-
nored in spite of its importance, particularly in diesel engines. The reason for using empirical cor-
relations is due to the complex nature of heat transfer during combustion and expansion stages.
The complexity arises from the large number of parameters inuencing the process such as the
thermo-physical properties of the products (viscosity, density, thermal conductivity, coecient
of heat transfer, specic heat, emissivity), their physical state (pressure, temperature, velocity rel-
ative to the surfaces of the combustion chamber) and eect of the reacting uid in turbulent mo-
tion and cyclic nature of the process. The best known empirical correlations [2125] estimate the
heat losses in the form of Newtons law of cooling with the overall heat transfer coecient made
up from convective and radiative components. The convective component is usually derived from
the well known Nusselt number correlation,
Nu aRe
b
; 20
where a and b are engine- and test-specic values. The radiative component is usually based on the
StefanBoltzmann equation for black body radiation in the form,
h
r
c
T
4
g
T
4
w
T
g
T
w
; 21
where, again, c is a engine- and test-specic constant. More realistic models for the radiative com-
ponent of heat transfer in engines were proposed which accounted for the eect of the radiation
properties of the gases [26,27]. However, estimates for of the radiative component from dierent
studies vary signicantly putting the ratio of radiation to the overall heat transfer anywhere be-
tween 0% and 50% [27]. The situation with the convective component is not much better as dif-
ferent correlations yield vastly dierent values for the coecient of heat transfer. Fig. 5 shows the
convection heat transfer coecient computed for the test engine using four correlations. The max-
imum value of h
c
ranges between 1500 and 7000 W/m
2
K. Woschni [23] compared several heat
transfer correlations with his own and found that the maximum values range between 1500
and 4500 W/m
2
K. The large variations in the values of the convection heat transfer coecient
10 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
predicted by dierent models and the fact that the radiation heat transfer component is often ne-
glected makes the selection of an appropriate model for the heat losses problematic.
In the current investigation, in the absence of a universally applicable heat transfer model for
diesel engines, a simple method proposed by Wiebe [28] is used as a rst approximation. This
method assumes that the cumulative apparent heat release makes up a constant fraction n of
the fraction fuel burned. The latter is then given by
v
t

v
i
n
: 22
Eq. (22) requires the knowledge of the coecient of combustion eectiveness n. This factor is
assumed at the start of computations then modied by successive iterations until the fraction fuel
burned approaches a value of 1.0. This equation can be used up to the moment when v
imax
is
reached.
The losses can then be expressed as
v
L

1
n
1
_ _
v
i
: 23
These losses are shown graphically in Fig. 4.
3. Results and discussion
Fig. 6 shows the results of sample calculations comprising the measured pressure trace, calcu-
lated average gas temperature and heat release characteristics. Two settings of the fuel injection
control system were used to produce a torque of 150 N m at 1800 rpm with both fuels. This
Fig. 5. Heat transfer coecients from dierent correlations calculated for the test engine.
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 11
ARTICLE IN PRESS
was feasible because the engine was operating well below its rated power. Were the engine in-
tended to produce the same power as the standard conguration, high capacity injectors and, pos-
sibly, a larger fuel injection pump would have been required. Under the current test regime, the
engine required 28.5 mg of diesel fuel and 35.86 mg of DOE per cycle, an increase of almost 26%.
This is an indication of the amount of extra fuel mass that needs to be carried by a vehicle or lo-
comotive for the same distance travelled when using diesel fuel. The decrease in pressure and tem-
perature with the DOE is due to the shift of the combustion process to the expansion stroke as a
result of retarded start of combustion (Fig. 6a) and quenching of the combustion process by the
water in the fuel. The role of water in the onset of micro-explosions of the DOE fuel droplets on
the combustion process also needs to be considered [29]. This decrease in temperature is the main
reason for the reduction in NO
x
emissions with emulsied fuels. The fuel burning rate during the
premixed phase is lower for the DOE because of the reduced amount of diesel fuel being injected
during the ignition delay phase. However, the reduction in the burning rate is not as great as was
expected due, possibly, to the higher fuel injection rate, which improves the air fuel mixing process
and accelerates the combustion process. The heat release characteristics also show that the burn-
ing rate during the diusion phase is almost the same for both fuels despite the late start of com-
bustion of the DOE. This is why the heat-to-work conversion eciencies (both indicated and
brake) remain almost unchanged when the engine is switched to DOE.
Fig. 6. Pressure, temperature and heat release characteristics of the combustion of standard diesel fuel and DOE at
150 N m and 1800 rpm.
12 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS
4. Conclusions
A relatively simple single-zone model has been developed to calculate heat release characteris-
tics in internal combustion engines using diesel oil emulsions and standard diesel fuel. The known
sensitivity in this type of calculation to the ratio of specic heats c
p
required use of calculated val-
ues for the two fuels tested based on the assumption of xed composition of the combustion prod-
ucts (no dissociation). These equations yield values of c
p
that better reect real processes in
engines. The model is a suitable tool for quick evaluation and interpretation of the performance
of dierent engines with dierent congurations or fuels and for the same engine under variable
operating conditions. It is also useful when used to monitor real-time engine heat release charac-
teristics for diagnostic purposes. While included in the model, a more rigorous treatment of heat
transfer losses needs to be developed for the calculation of fraction fuel burned before this preli-
minary treatment can be conrmed. Further work is planned to develop and validate a more real-
istic heat transfer module that can account for the eect of the properties of the combustion
products on the radiative component of the losses.
Acknowledgements
This work is supported by a grant from the Rail CRC. The authors wish to thank Mr. Khaled
Al-Khalee for providing the experimental engine test data.
References
[1] J.E. Nicholls, I.A. El-Messiri, H.K. Newhall, Inlet manifold water injection for control of nitrogen oxidestheory
and experiment, SAE Paper No. 690018.
[2] S.J. Lestz., R.B. Melton Jr., E.J. Rambi, Feasibility of cooling diesel engines by introducing water into the
combustion chamber, SAE Paper No. 750129.
[3] R.J. Crookes, M.A.A. Nazha, M.S. Janota, T. Storey, Investigation into the combustion of water/diesel fuel
emulsions, SAE Paper No. 800094.
[4] G.E. Andrews, K.D. Bartle, S.W. Pang, A.M. Nurein, P.T. Williams, The reduction in diesel particulate emissions
using emulsied fuels, SAE Paper No. 880348.
[5] M. Tsukahara, Y. Yoshimoto, T. Murayama, W/O emulsion realizes low smoke and ecient operation of DI
engines without high pressure injection, SAE Paper No. 890449.
[6] M. Ishida, Z. Chen, An analysis of the added water eect of NO formation in DI diesel engines, SAE Paper No.
941691.
[7] Z. Chen, M. Ishida, An analysis of the added water eect on NO formation in DI diesel engines, SAE Paper No.
941691.
[8] Z. Li, H. Sano, M. Tsukahara, Y. Yoshimoto, NO
x
reduction with EGR in a diesel engine using emulsied fuel,
SAE Paper No. 982490.
[9] Comparative Analysis of Vehicle Emission Using PuriNOx Fuel and Diesel Fuels, Air Improvement Resource,
Inc., April 4, 2001.
[10] EPA, Impacts of Lubrizols PuriNox water diesel emulsion on exhaust from heavy duty engines, Draft Report,
EPA20-P-02-007, December 2002.
[11] C.-Y. Lin, K.-H. Wang, Diesel engine performance and emission characteristics using three-phase emulsion as fuel,
Fuel 83 (2004) 537545.
J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx 13
ARTICLE IN PRESS
[12] G.H. Abd Alla, A.A. Soliman, O.A. Badar, M.F. Adb Rabbo, Combustion quasi-two zone predictive model for
dual fuel engines, Energy Conversion and Management 42 (2001) 14771498.
[13] P.A. Lakshminarayanan, Y.V. Aghav, A.D. Dani, P.S. Mehta, Accurate prediction of the heat release in a modern
direct injection diesel engine, Proceedings of the Institute of Mechanical Engineers 216 (2002) 663675.
[14] M. Lapuerata, O. Armas, V. Bermudez, Sensitivity of diesel thermodynamic cycle calculation to measurement
errors and estimated parameters, Applied Thermal Engineering 20 (2000) 843861.
[15] M. Lapuerata, J.J. Hernandez, O. Armas, Kinetic modelling of gaseous emissions in a diesel engine, SAE Technical
Paper 2000-01-2939, 2000.
[16] J.I. Ghojel, Computer time reduction in rst law analysis of combustion systems, in: Proceedings International
Conference on Fluid and Thermal Energy Conversion94, vol. 1, 1994, pp. 335342 (ISSN 0854-9346).
[17] I.I. Wiebe, M.F. Farafanov, V.B. Shaliganova, A method for the investigation of ICE cycles using Ural-2 computer
(in Russian), automobiles, tractors and engines, ChPI Collected Scientic Papers, No. 52, 1969, pp. 232250.
[18] G. Hohenberg, Experimentelle Erfassung der Wandwarme in Kolbenmotoren (Experimental Acquisition of the
Wall Heat in Piston Engines), Habilitation Thesis, Technical University of Graz, 1980.
[19] V. Arkhangelsky, M. Khovakh, Y. Stepanov, V. Trusov, M. Vikhert, A. Voinov, Motor Vehicle Engines, Mir
Publisher, Moscow, 1971, 619 p.
[20] G.J. Van Wylen, R.E. Sonntag, Fundamentals of Classical Thermodynamics, Wiley, New York, 1978.
[21] G. Eichelberg, Some new investigations on old combustion engine problems, Engineering 148 (1939) 463547.
[22] W.J.D. Annand, Heat transfer in the cylinders of reciprocating internal combustion engines, Proceedings of the
Institution of Mechanical Engineers 177 (36) (1963) 973990.
[23] G. Wochni, A universally applicable equation for the instantaneous heat transfer coecient in the internal
combustion engine, SAE Transactions 76 (1967) 30653083.
[24] G.F. Hohenberg, Advanced approaches for heat transfer calculations, SAE Paper 790825, SP-449, 1979.
[25] N.Kh. Dychenko, et al., Theory of Internal Combustion Engines, Mashinostroyeniye, Leningrad, 1974.
[26] T.J. Callahan, T.W. Ryan, Acquisition and interpretation of diesel engine heat release data, SAE Paper 852068,
1985.
[27] D.N. Assanis, J.B. Heywood, Development and use of computer simulation of the turbocharged diesel system for
engine performance and component heat transfer studies, SAE Paper 860329, 1986.
[28] I.I. Wiebe, New Insight into the Engine Cycle, Mashgiz, Moscow, 1962 (Chapter 2), pp. 3589 (in Russian).
[29] J.C. Lasheras, A.C. Fernandez-Pello, F.L. Dryer, Initial observations on the free droplet combution of water-in-
fuel emulsions, Combustion, Science and Technology 21 (1979) 114.
14 J. Ghojel, D. Honnery / Applied Thermal Engineering xxx (2005) xxxxxx
ARTICLE IN PRESS

Вам также может понравиться